Thoughts on the ocean, the environment, the universe and everything from nearly a mile high.

Panorama of The Grand Tetons From the top of Table Mountain, Wyoming © Alan Holyoak, 2011

Monday, November 5, 2012

Understanding Climate Change Part 3 - Earth's Energy Budget - Radiation From Sun and Earth

Earths’ Energy Budget: Radiation From the Sun and Earth

(Reading #3 for my course on Climate Change, Alan Holyoak, PhD)

Daily Objectives
  1. Be able to explain what radiation is, including the difference between direct solar radiation and infrared radiation.
  2. Be able to identify the wavelengths of radiation that are important to global climate.
  3. Be able to explain why there is a latitudinal difference in the amount of energy received each year.
  4. Be able to explain what the Earth’s energy budget is, and the factors that affect the flow of energy into and out of the atmosphere.
  5. Be able to list important greenhouse gases and explain the term `global warming potential’.
Introduction

The climate of the Earth is affected by the rate of gain and loss of energy.  This dynamic exchange of energy entering and leaving Earth is used to describe its energy budget.  Solar radiation is the main form of energy that is important to climate.  While the sun is the major source of radiant energy for the Earth, the Earth, the atmosphere, and everything warmer than absolute zero also emit radiation at all times and in all directions.  It is therefore important to understand what radiation is, as well as the processes involved in energy gains and losses at Earth’s surface in order better understand climate change.

About Radiation

            Radiation is energy, and is therefore subject to the Laws of Thermodynamics.  Since these laws apply, it is important to remember that energy cannot be created or destroyed, but it can change forms, and every time it changes form entropy increases and is released as waste heat.  Energy is released by everything with a temperature above absolute zero (Fig. 1).  The temperature of an object determines the wavelength of energy it releases.   Absolute zero is the lowest temperature that can theoretically exist, and is indicated by 0o Kelvin, -273oC, or -460oF.


Figure 1. Thermal image approximating what we might look like to each other if we could see infrared radiation. (Image courtesy of Wikimedia Commons.)

             Think of radiation as a three-dimensional wave of energy.  These waves travel at the speed of light, which is 3x108 meters/second or 186,000 miles/second.  Note that any wave has a particular wavelength or distance from crest to crest (Fig. 2).  The units for a wavelength of radiation are usually micrometers (a millionth of a meter and designated by “μm”) or nanometers (a billionth of a meter and designated by “nm”).
Figure 2.  Each spectrum of radiation has its own unique wavelength.  Spectra with longer wavelengths (upper image) carry less energy than those with shorter wavelengths (lower image). (Image courtesy of Dr. Hipps, Utah State University.)


Wavelengths of Radiation

            The range of wavelengths of electromagnetic radiation in the universe is enormous.  Figure 3 shows this range and indicates wavelengths of radiation important to Earth’s climate.  Many people are surprised when they learn that visible radiation or “light” comprises only a tiny segment of the total range of the electromagnetic spectrum of radiation.  Your eyes sense only this narrow range of wavelengths.  Even though there is an enormous range of electromagnetic radiation, only small section of the range is important to Earth’s energy budget.


Figure 3. The electromagnetic spectrum of radiation.  Wavelengths of the spectrum of particular interest to global climate include ultraviolet (UV) radiation of ~10-380nm, visible light from ~380-740nm, and infrared (IR) radiation of ~740nm-300μm. (Image courtesy of Wikimedia Commons.)

            The amount of radiation emitted and the wavelengths of that emission are determined by the temperature of an object.  Table 1 shows the key principles of radiation emission, wavelength distribution, and temperature of emissions from the sun and Earth.

Table 1.  Characteristics of radiation released by the sun and Earth.

Emitting body
Temperature
Radiation characteristics
Sun
6000 K (~5700oC)
Larger emission at shorter wavelengths
Earth
   290 K (~17oC)
Smaller emission at longer wavelengths

Radiation Emission of Sun and Earth

            Since the sun and Earth have significantly different temperatures we expect their radiation emissions to be quite different.  The sun is about 6000oK (5730oC or 10,340oF).  The average temperature on Earth is about 290oK (17oC or 62oF).  There are therefore large differences in both the intensity and wavelengths emitted by these two objects. Visible light comprises about half the total emission from the sun (Fig. 4), while the vast majority of radiation released by the much cooler Earth is in the range of infrared radiation. 

 
Figure 4. The sun emits mainly high-energy, short-wavelength radiation, including the visible spectrum of light, and the Earth emits mainly low-energy, long-wavelength radiation, mostly in the form of infrared radiation or heat. (Image courtesy of Dr. Hipps, Utah State University.)

Some important points to know

1.     The size of emission and distribution of wavelengths are related to temperature.  The sun emits more energy and in shorter wavelengths than the Earth.  The Earth, which is cooler than the sun, emits less energy and at longer wavelengths.

2.     About half of solar emission lies in the wavelengths of visible light (380-740nm), and it peaks near 500nm.

3.     Earth emissions cover a large range of wavelengths, and peak at about 10 μm.  And because infrared radiation falls outside of the range of visible light, we cannot see it.


Total Radiation Emission For All Wavelengths

            The total sum of the emission of radiation over all wavelengths can be calculated from a law of physics in which the energy emitted is calculated by multiplying a constant “k” that includes the ability of an object to absorb energy that is a value between 0 and 1, the area of the object, and Stephan’s constant (don’t get hung up on this) by the temperature “T” of the object in degrees Kelvin taken to the 4th power as shown in the this equation:

Emission = k * T4

This means that when we know the temperature of an object we can readily calculate the total radiation (energy) it emits.  Because emission changes with temperature to the fourth power, small changes in temperature translate into large changes in emission!  The ability to make these kinds of measurements and carry out these calculations makes it possible for climatologists to track the flow of energy throughout Earth’s climate system.

Units of Energy

            Since radiation is energy, we should use the appropriate units to refer to it: Joules.  A Joule is a unit of energy or work.  Maybe this will help you understand how much energy is in one Joule.  Imagine that there is an average sized apple sitting on the ground.  You expend about one Joule of energy when you lift that apple to a height of one meter at a constant speed.  Most of the time, though, we refer to energy in terms of Watts.  Does this sound more familiar?  Going back to our example of lifting an apple, you expend about one Watt of energy when you lift the apple to the height of one meter in one second.  So if you lifted 100 apples at the rate of one meter/second you’d expend 100 Watts of energy.  Though Joules are the best unit for referring to energy, Watts are used when we want to measure the rate at which energy is moved or used.  The units most conveniently used to refer to radiant energy are therefore Watts per meter squared.  This lets us know how much energy we are talking about in terms of an area of interest.  So don’t be surprised to see various values in Watts and Watts m-2 throughout the rest of the semester.  You therefore need to become familiar and comfortable with this unit.  By the way, if you have ever ridden a bicycle that has a friction generator used to power a light, you can really feel it when you ride!  Most of these generators produce only a few Watts of power (Fig. 5).

Figure 5. A friction generator used to provide power for bicycle head and tail lights (3 Watts).
(Image courtesy of Amazon.com)

Seasonal Changes in Solar Energy Received on Earth

            You already know that the tilt of the Earth produces our seasons, but here are some additional things that were not covered in that reading that you need to be aware of.
            As a result of Earth’s axial tilt the sun’s rays can be directly overhead at noon only between latitudes of 23.5oN (the Tropic of Cancer) and 23.5oS (the Tropic of Capricorn).  The latitude where the sun is directly overhead at noon moves north and south with the seasons.  At all other places on the planet the elevation of the sun from the horizon can never reach vertical or 90o overhead.
            On exactly two days a year the sun is vertical at noon directly at the equator.  These events are called equinoxes.  The autumnal equinox is about 22 September, and the vernal equinox is about 22 March.  Let’s follow the movement of the sun’s location over a year. 
After the autumnal equinox the location of the sun’s vertical rays moves gradually south until about 22 December when it is vertical at the Tropic of Capricorn.  This day is the winter solstice.  After this date the latitude where the sun is vertical moves north until it reaches the equator at about 22 March.  This is the vernal equinox.  The location where the sun’s rays are vertical at noon continues to move north until it reaches the Tropic of Cancer on about 22 June.  This is the summer solstice.  The location where the sun is vertical at noon then moves south until it reaches the equator near 22 September on the autumnal equinox (Table 2).

Table 2. Summary of dates important to seasonal change on Earth.
Event
Date
Latitude sun is vertical at noon
Autumnal equinox
22 September
Equator
Winter solstice
22 December
Tropic of Capricorn
Vernal equinox
22 March
Equator
Summer solstice
22 June
Tropic of Cancer

            Remember that 50% of the Earth’s surface is always being illuminated.  Even so, because the sun is vertical above the equator at noon only on the equinoxes, on only those dates the day length is 12 hours everywhere on Earth.
            Because of the tilt of Earth’s axis, polar regions of the Earth have periods of complete darkness and others that have constant light at various times of the year.  This results in two more latitudes that are special when we think about seasons and climate.  On the summer solstice no solar radiation reaches farther south than 90 minus 23.5o, or 66.5o S.  This latitude is the Antarctic Circle.  So at any latitude greater than 66.5o S there is a period of time when it receives no solar radiation at all.  Conversely, on the winter solstice the region of complete light reaches only as far north as 66.5o S. Similarly on the summer solstice, the region of 24-hour daylight reaches as far south as 66.5o N.  This latitude is called the Arctic Circle.  Within the Antarctic Circle there is one day per year of no light, 22 June, and one day per year of 24-hour light, 22 December.  As you move toward the South Pole from 66.5o S or the North Pole from 66.5o N, the number of days of darkness or light per day grows until at the North and South Poles 6 months pass when the sun never rises above the horizon, and 6 months of when the sun doesn’t drop below the horizon.
            The rate of change of the sun angle overhead varies with the seasons.  The position of the sun in the sky changes most slowly at the solstices and fastest at the equinoxes.  You may have noticed that our day length changes slowly during the middle of the summer and the middle of the winter, and that the duration of day length changes much faster during the fall and spring.  Near the equator there is little seasonal change in sun angle or day length.  The day length in the tropics is therefore always about 12 hours.
            There are also seasonal changes in the path the sun takes across the sky, especially the farther you go from the equator.  The sun will rise from exactly the east and set exactly in the west only on equinoxes.  For example, let’s consider the track of the sun across the northern hemisphere between 22 September and 22 March.  On each day the sun will rise south of true east and set south of true west.  The maximum elevation reached by the sun at local noon will reduce every day until it reaches its lowest height on 22 December – the shortest day of the year in the northern hemisphere.  So during the winter not only is the height of the sun in the sky lower, but the length of the arc it makes across the sky is also shorter.  Both of these factors result in less total solar radiation being received by the surface during the fall and winter months in the northern hemisphere than during the Spring Summer.  During the northern hemisphere summer the sun rises north of true east and sets north of true west.  The arc of the sun’s path across the sky is now much longer than in the Fall/Winter.  In addition, the maximum elevation of the sun reached at noon also increases.  So total solar radiation is greater in the summer than in the winter. 
            What is the largest elevation angle the sun can reach at noon at any given latitude?  You can calculate this by subtracting 23.5o from the latitude of interest, and then subtracting this difference from 90o.  For example, the latitude of Rexburg is 43.8oN.  The maximum solar elevation at this latitude is: 90 – (43.8 – 23.5) = about 69.7o.  So, on 22 June the sun will reach its maximum elevation at noon, and will be at an angle of 69.7o overhead. 
            Why bother going into all of this?  Latitudinal variations in solar radiation have a massive effect on climate!

Earth’s Energy Budget

            Earth’s energy budget is the dynamic balance between total energy received and total energy emitted.  The atmosphere affects the rate of movement of radiation as it passes through it so calculating Earth’s energy budget can be challenging.  One very important thing to understand is that all energy received is eventually emitted back into space. 
We will start by looking at the rate at which solar radiation enters and infrared radiation leaves.  Figure 6 shows that 100% of the energy entering Earth’s system eventually leaves and goes back into space.  The relative rates at which energy enters, the length of time it stays in the atmosphere, and the length of time it takes energy to leave determine whether Earth is warming or cooling (Table 3). 

Table 3. Effects of the rates of radiation entering and leaving Earth’s atmosphere.
Differential rate of energy flow
Average global temperature
Energy entering  > Energy leaving
Warming
Energy entering = Energy leaving
Unchanging
Energy entering < Energy leaving
Cooling

            Radiation can have many fates as it passes through the atmosphere.  Some of it can pass all the way through the atmosphere and strike the Earth’s surface. However, radiation can be also absorbed, reflected, or scattered by the atmosphere.  It is also absorbed or reflected by the surface.

Fates of Solar Radiation

            Solar radiation entering Earth’s atmosphere has a variety of fates.  Solar radiation can be absorbed, reflected, or scattered by the atmosphere before it reaches the surface (Figs. 6 & 7).  The remaining solar radiation that passes all the way through the atmosphere and strikes the Earth’s surface and called direct solar radiation; this is the light you see when you glance up at the sun.  Direct solar radiation can be absorbed or reflected by Earth’s surface.  

Figure 6.  Simplified view of Earth’s energy budget. Yellow arrows represent solar radiation.  Orange arrows represent solar energy that has been absorbed.  Red arrows represent the release of absorbed energy as infrared radiation. (Image courtesy of NASA.)

Figure 7. Fates and pathways of incident solar radiation, including absorption, reflection, and scattering. (Image courtesy of Dr. Hipps, Utah State University.)
Absorption

            The main factor that determines the rate at which energy moves through the atmosphere is the composition of the atmosphere itself.  Some gases can absorb certain wavelengths of solar radiation.  Most of this absorption occurs in the ultraviolet (UV) region.  Most UV radiation is absorbed high in the stratosphere.  Major players involved in this absorption are ozone, oxygen, and N2O (nitrous oxide).  Some UV radiation reaches the Earth’s surface as part of direct solar radiation, and this is what causes sunburns and skin cancers.
            Small particles called aerosols and clouds can absorb and reflect radiation.  Clouds are highly reflective to solar radiation.  Interestingly, the atmosphere does not absorb much energy in the visible wavelengths of light (380-740nm).  Solar radiation absorbed in the atmosphere results in an increase in the temperature of the absorbing gases. 

Scattering

            When solar radiation approaches some small particles or air molecules, a complex interaction takes place that causes the incident beam (the ray of solar energy) to be redirected in random directions.  This is scattering.  Scattered light can be scattered again and again, depending on its path and what it encounters.  Ultimately, this redirected light finds its way back into space or to the Earth’s surface.  Scattered light that reaches the surface in this way is called diffuse radiation, and thus differs from direct radiation that reaches the surface without interference.  On a cloudy day, and at times just before sunrise and just after sunset, all radiation is diffuse.  Whenever direct sunlight is present, a combination of direct and diffuse radiation reaches the surface.
            Particles and air molecules smaller than the wavelengths of visible light cause something called Rayleigh scattering.  Causes shorter wavelengths of visible light (greens, blues, and violets) to be scattered more than longer wavelengths of visible light (reds, oranges, and yellows).  As a result when the sun is more directly overhead the sky looks blue.  When the light has to pass through more of the atmosphere before it reaches us, as at dawn and dusk, shorter wavelengths are all scattered, leaving only the reds and oranges, and this can give us colorful sunrises and sunsets. 
            Since scattering redirects radiation into all directions, some of radiation is always reflected into space.  So, scattering always reduces the total solar radiation reaching the surface.

Surface Reflection

            The reflectivity of a surface refers to the proportion of solar radiation reflected by the surface in relation to the total amount of radiation reaching the surface. This factor is called the albedo.  Every object will have an albedo between 0.0 and 1.0.  An albedo of 0.0 means that a surface is perfectly black and absorbs all wavelengths of radiation; none is reflected.  An albedo of 1.0, on the other hand, means that a surface is perfectly white and all wavelengths of radiation are reflected.  The albedo of objects in the atmosphere and on Earths’ surface is critical in the climate system.
            When the entire Earth-atmosphere system is considered, Earth’s average albedo is about 0.3.  So about 30% of the incident solar radiation is reflected back into space (see Fig. 6).  This 30% includes reflection off of Earth’s atmosphere, clouds, and surface.

Infrared Radiation and the Atmosphere

            The Earth and molecules in the atmosphere absorb high-energy solar radiation, and then emit longer wavelength, low-energy infrared radiation (IR).  This IR eventually makes its way back out into space.  The rate at which IR makes its what through the atmosphere, however, is determined mainly by the following factors:
  • IR is absorbed by: liquid water in clouds, water vapor, carbon dioxide (CO2), methane (CH4), nitrous oxide (N2O), ozone (O3), chlorofluorocarbons (CFCs), and other greenhouse gases.
  • IR absorbed by atmospheric components increases atmosphere retention and then emission of IR in all directions
  • This non-directional emission of IR slows the rate of IR loss to space

The Greenhouse Effect

The absorption of some of the IR from the Earth by the atmosphere has a major effect on climate.  The IR energy absorbed by the atmosphere increases the overall energy content of the atmosphere.  This causes an increase in the IR emission by the absorbing gas or cloud.  However, the IR travels in all directions when it is emitted.  So, some of the IR energy ends up moving out into space, but some finds its way back toward the surface.  These IR absorbing molecules act to reduce the net rate of loss of energy from the Earth to space.  This process is the Greenhouse Effect.  The gases that perform this absorption in the atmosphere are called Greenhouse Gases.
Greenhouse gases keep the Earth warmer that it would otherwise be.  When the amount of greenhouse gases in the atmosphere increases, the rate of loss of IR into space slows down and atmospheric average temperatures rise.  The presence of greenhouse gases and clouds maintain the Earth at a much warmer temperature that it would otherwise have.  Earth’s average temperature is about 17oC or 63oF, but calculations have shown that without greenhouse gases the Earth would have an average temperature of about -18oC or 0oF! 
At this point it is important to note that not all greenhouse gases behave the same way in the atmosphere.  Water vapor, for example, is a strong greenhouse gas, but it does not accumulate indefinitely in the atmosphere.  Once the atmosphere reaches its full capacity of water vapor and the air cools even slightly, water vapor will condense into water droplets and fall as precipitation.  Other kinds of greenhouse gases such as carbon dioxide, however, can accumulate more or less indefinitely in the atmosphere, allowing their effects to increase over time.

Fates of Infrared Radiation (IR)

Figure 8 shows the pathways of IR in the Earth-atmosphere system.  The major processes that affect IR radiation balance are also displayed.  Recall that the surface and atmosphere both emit energy.  The source of IR is mainly from the Earth surface.  The surface is heated by solar radiation, and later releases that energy when it emits IR into the atmosphere.  The amount of IR emitted by the surface is proportional to surface temperature.
There is significant absorption of IR by water in clouds, water vapor, and other greenhouse gases once the Earth releases the IR.  Because the air releases IR in all directions, a significant proportion of IR ends up returning back to the surface.
Clouds play a major role in the climate system because they contain droplets of liquid water.  Clouds reflect incident solar radiation and reduce the amount of direct solar radiation that reaches the surface.  However, clouds are also strong absorbers of IR, which slows the loss of energy from the surface.  The combined effects of clouds on the total budget of solar radiation and IR are therefore complex.  A big issue at present is whether changes in clouds due to increased atmospheric temperature will act to accelerate of diminish further warming since they simultaneously absorb IR energy and reflect incident solar energy back into space.
There is a strong relationship between IR balance and temperature.  For example, an increase in the concentration of greenhouse gases results in imbalance between energy received and energy emitted into space.  The average temperature would consequently increase because there would be a smaller IR loss.  But the higher temperature would also produce increases in IR surface emission and increase IR loss into space.  Eventually IR loss would reach a level where total losses would again balance solar input.  At that point a new equilibrium is reached, but at a higher average global temperature.

Figure 8.  Fates of infrared radiation.  (Image courtesy of Dr. Hipps, Utah State University.)

More about Greenhouse Gases

            Direct measurements of the atmosphere have revealed that human activities are increasing the concentrations of some greenhouse gases in the atmosphere, and there is reason for concern about their effects on the global climate (Fig. 9). For example, atmospheric scientists examined all known natural and anthropogenic (human-produced) sources of CO2 and concluded that virtually all recent increases in CO2 is attributable to two main sources: 1) burning fossil fuels for energy, transportation, and industry, and 2) cement production.  This trend is exacerbated by the destruction or degradation of key terrestrial ecosystems, such as forests, that have the potential to capture and store excess carbon emissions.  Methane has also been steadily rising during the same time period as CO2.  The cause of methane rise is less well documented than for CO2, but it also appears to be linked to human activities such as draining wetlands and swamps, increasing cattle production, and other activities.  Increases in N2O are related to fertilizer use, combustion, and industry.


Figure 9. Trends of greenhouse gas concentrations in the atmosphere, 1978-2011. (Image courtesy of NOAA.)

           Chlorofluorocarbons (CFCs) are experiencing an interesting trend of concentrations in the atmosphere (Fig. 9).  There was a rapid increase in CFC concentrations through the mid-1990s.  Since that time, however, CFC concentrations appear to have leveled off.  This is not surprising, since the Montreal Protocol, an international agreement reached in 1987 banning the production and distribution of CFCs went into effect in 1989.  This agreement was formalized after governments recognized overwhelming scientific evidence showing that CFCs are the primary agent destroying the stratospheric ozone layer. 
CFCs were used initially as a replacement for ammonium as a refrigerant.  CFCs have several advantages over ammonia.  For example, a refrigeration system containing ammonia could leak, and the ammonia readily evaporates.  The problem with this is that ammonia is a poisonous gas.  How would it be to be killed by your refrigerator!?  CFCs, on the other hand, are excellent refrigerants and they are not toxic.  They were also believed to be chemically stable, and they are under most conditions, but when CFCs make their way into the stratosphere they are bombarded by UV radiation.  This causes CFCs break down and release individual atoms of chlorine.  These chlorine atoms react with ozone, splitting one atom of oxygen away and forming chlorine monoxide (ClO).   The ClO then also breaks apart in the presence of UV light, and the chlorine ion then breaks down another atom of ozone, and so on and so son.  It is estimated that a single chlorine atom in the stratosphere will break down as many as 100,000 molecules of ozone before being bound up in a longer-lived molecule.  This is why ozone depletion occurs in the presence of CFCs.  This CFC-driven reduction of the stratospheric ozone layer allowed increased amounts of UV radiation to reach the surface.  UV radiation can cause biological problems ranging from reduced photosynthetic ability in plants to skin cancers and eye cataracts in animals.  And, CFCs are powerful greenhouse gases.
You may be wondering why there are still CFCs in the atmosphere if they were banned as of 1989.  CFCs are long-lived molecules (e.g., CFC-11 = 45 yrs, CFC-12 = 100 yrs), and there are still a lot of CFCs still in people’s refrigerators, air conditioners, etc., which continue to leak and be released into the atmosphere.
Anyway, the range of wavelengths of energy and amount of energy individual molecules can absorb varies from one kind of greenhouse gas to another.  Since this is the case, we need to find a way we can compare the effects of different greenhouse gases to each other in an apples-to-apples type of comparison.  This is where the global warming potential (GWP) of greenhouse gases helps out.  Table 4 shows the GWP of key greenhouse gases in relation to CO2, which has been assigned a baseline GWP of 1.0. 
            Note the large differences in the effect of different greenhouse gases on temperature.  Compared to CO2, a single molecule of methane has a 21x greater GWP.  But, also notice the concentrations!  The value for CO2 is almost 200x more abundant in the atmosphere than methane, and about 1 million times more abundant than CFC-12.  This makes CO2 is the most important of the greenhouse gases when you consider its GWP and its overall concentration in the atmosphere, and this is the greenhouse gas humans are currently emitting into the atmosphere at the greatest rate. 

Table 4. Global warming potential (GWP) of selected greenhouse gases.
Greenhouse gas
Atmospheric concentration
GWP
CO2
~390 pp million
1
CH4
~     2 pp million
21
N2O
~ 320 pp billion
310
CFC-12
~540 pp trillion
8500
HCFC-22
~100 pp trillion
1350

Net Radiation

            The value of radiation that concerns us when we consider Earth’s energy budget is the actual amount of radiant energy available at the surface, called net radiation (Rnet).  This is calculated by accounting for all solar and infrared radiation as follows:

Rnet = (Incident Solar – Reflected Solar) + (Incident IR – Surface IR)

            The sum of these radiation streams represents the amount of energy available to drive Earth’s climate.  OK, what is this energy used for?
            About 77% of Rnet is used to evaporate water, and the remaining 23% is used to heat the surface.  In other words, the ocean absorbs the vast majority of solar energy.  These results are not surprising when we consider that about 71% of Earth’s surface is covered by water.  The large amount of energy associated with the evaporation process means that this factor has large effects on global and regional climate systems.
            If the temperature of Earth were constant, then the Rnet, when averaged over the entire planet for an entire year, would equal zero.  Sadly, such is not the case.  Earth currently has a positive Rnet.

Patterns of Radiation Over the Earth

Because of the tilt of the Earth and its orbit around the sun, there is a large seasonal change in radiation over the surface.  There is also a large variation in received and emitted radiation with latitude.  Figure 10 shows the annual distribution of solar energy received over an entire year by latitude.  It is not a surprise that the tropics receive the greatest amount of solar radiation and has the highest Rnet over a year.  These values decrease with distance from the equator until a minimum is reached at the poles.

Figure 10. Total amount of solar radiation received by latitude. (Image courtesy of Dr. Hipps, Utah State University.)

It is useful to look at the net value of absorbed minus emitted radiation for each latitude band, over a year.  This requires plotting both the solar radiation received and IR lost by the surface.  Figure 11 shows this difference by latitude.  In the tropics there is more energy gained at the surface than is lost every year.  At higher latitudes the opposite is observed where more IR is lost than the amount of solar energy gained.  So there is a continuous deficit of energy at higher latitudes.  The regions of surplus change to deficits at about 37o north and south.

Figure 11. Annual net radiation by latitude.  (Image courtesy of the Univ. of Wisconsin.)


If this unequal distribution of radiation by latitude were the only process that affects climate, what would Earth’s temperature profile look like?  For one thing, the tropics would not just be hot, they would be unbearably hot, and the rest of the planet would be either cold or unbearably cold.  This is not observed because heat is constantly transported away from the tropics to higher latitudes via air and ocean currents.
            The unequal distribution of radiation over the Earth is the most important process of the global climate system.  The physical processes of the atmosphere and ocean are continually working to transfer energy from the tropics towards the poles.  These massive transports of heat have a massive impact on global climate. 
It is useful to look at how the distribution of this energy varies by season.  Figure 12 is an image showing net radiation over the Earth in January and in June.  Red and orange colors represent positive net radiation values, and blue shows negative values.  Remember that the sun is near vertical at 23.5oS during the northern hemisphere winter.

Figure 12. Intensity of net radiation during January (left) and June (right).  (Images courtesy of the Univ. of Wisconsin.)

In the next few readings you will learn about processes that transport energy (heat) around the planet.

Source material
Hipps, LE. 2010. Personal communication and readings produced by Dr. Hipps. Professor of Atmospheric Science, Department of Plants, Soils, and Climate.  Utah State University.

No comments:

Post a Comment